UNITARY REPRESENTATIONS and COMPLEX ANALYSIS David A
Total Page:16
File Type:pdf, Size:1020Kb
UNITARY REPRESENTATIONS AND COMPLEX ANALYSIS David A. Vogan, Jr. Department of Mathematics Massachusetts Institute of Technology Contents 1. Introduction 2. Compact groups and the Borel-Weil theorem 3. Examples for SL(2; R) 4. Harish-Chandra modules and globalization 5. Real parabolic induction and the globalization functors 6. Examples of complex homogeneous spaces 7. Dolbeault cohomology and maximal globalizations 8. Compact supports and minimal globalizations 9. Invariant bilinear forms and maps between representations 10. Open questions 1. Introduction Much of what I will say depends on analogies between representation theory and linear algebra, so let me begin by recalling some ideas from linear algebra. One goal of linear algebra is to understand abstractly all possible linear transformations T of a vector space V . The simplest example of a linear transformation is multiplication by a scalar on a one-dimensional space. Spectral theory seeks to build more general transformations from this example. In the case of infinite-dimensional vector spaces, it is useful and interesting to introduce a topology on V , and to require that T be continuous. It often happens (as in the case when T is a differential operator acting on a space of functions) that there are many possible choices of V , and that choosing the right one for a particular problem can be subtle and important. One goal of representation theory is to understand abstractly all the possible ways that a group G can act by linear transformations on a vector space V . Exactly what this means depends on the context. For topological groups (like Lie groups), one is typically interested in continuous actions on topological vector spaces. Using ideas from the spectral theory of linear operators, it is sometimes possible (at least in nice cases) to build such representations from irreducible representations, which play the role of scalar operators on one-dimensional spaces in linear algebra. Here is a definition. 1991 Mathematics Subject Classification. Primary 22E46. Supported in part by NSF grant DMS-9721441 Typeset by AMS-TEX 1 2 DAVID A. VOGAN, JR. Definition 1.1. Suppose G is a topological group. A representation of G is a pair (π; V ) with V a complete locally convex topological vector space, and π a homomorphism from G to the group of invertible linear transformations of V . We assume that the map G × V ! V; (g; v) 7! π(g)v is continuous. An invariant subspace for π is a closed subspace W ⊂ V with the property that π(g)W ⊂ W for all g 2 G. The representation is said to be irreducible if there are exactly two invariant subspaces (namely V and 0). The flexibility in this definition—the fact that one does not require V to be a Hilbert space, or the operators π(g) to be unitary|is a very powerful technical tool, even if one is ultimately interested only in unitary representations. Here is one reason. There are several important classes of groups (including reductive Lie groups) for which the classification of irreducible unitary representations is still an open problem. One way to approach the problem (originating in the work of Harish- Chandra, and made precise by Knapp and Zuckerman in [KnH]) is to work with a larger class of \admissible" irreducible representations, for which a classification is available. The problem is then to identify the (unknown) unitary representations among the (known) admissible representations. Here is a formal statement. Problem 1.2. Given an irreducible representation (π; V ), is it possible to impose on V a Hilbert space structure making π a unitary representation? Roughly speak- ing, this question ought to have two parts. (1.2)(A) Does V carry a G-invariant Hermitian bilinear form h; iπ? Assuming that such a form exists, the second part is this. (1.2)(B) Is the form h; iπ positive definite? The goal of these notes is to look at some difficulties that arise when one tries to make this program precise, and to consider a possible path around them. The difficulties have their origin exactly in the flexibility of Definition 1.1. Typically we want to realize a representation of G on a space of functions. If G acts on a set X, then G acts on functions on X, by [π(g)f](x) = f(g−1 · x): The difficulty arises when we try to decide exactly which space of functions on X to consider. If G is a Lie group acting smoothly on a manifold X, then one can consider C(X) = continuous functions on X; Cc(X) = continuous functions with compact support; 1 Cc (X) = compactly supported smooth functions: C−∞(X) = distributions on X: UNITARY REPRESENTATIONS AND COMPLEX ANALYSIS 3 If there is a reasonable measure on X, then one gets various Banach spaces like Lp(X) (for 1 ≤ p < 1), and Sobolev spaces. Often one can impose various other kinds of growth conditions at infinity. All of these constructions give topological vector spaces of functions on X, and many of these spaces carry continuous repre- sentations of G. These representations will not be \equivalent" in any simple sense (involving isomorphisms of topological vector spaces); but to have a chance of get- ting a reasonable classification theorem for representations, one needs to identify them. When G is a reductive Lie group, Harish-Chandra found a notion of “infini- tesimal equivalence" that addresses these issues perfectly. Inside every irreducible representation V is a natural dense subspace VK , carrying an irreducible represen- tation of the Lie algebra of G. (Actually one needs for this an additional mild assumption on V , called \admissibility.") Infinitesimal equivalence of V and W means algebraic equivalence of VK and WK as Lie algebra representations. (Some details appear in section 4.) Definition 1.3. Suppose G is a reductive Lie group. The admissible dual of G is the set G of infinitesimal equivalence classes of irreducible admissible representa- tions of G. The unitary dual of G is the set G of unitary equivalence classes of b u irreducible unitary representations of G. b Harish-Chandra proved that each infinitesimal equivalence class of admissible irreducible representations contains at most one unitary equivalence class of irre- ducible unitary representations. That is, (1.4) Gu ⊂ G: This sounds like great news for the programb bdescribed in Problem 1.2. Even better, he showed that the representation (π; V ) is infinitesimally unitary if and only if the Lie algebra representation VK admits a positive-definite invariant Hermitian form h; iπ;K . The difficulty is this. Existence of a continuous G-invariant Hermitian form h; iπ on V implies the existence of h; iπ;K on VK ; but the converse is not true. Since VK is dense in V , there is at most one continuous extension of h; iπ;K to V , but the extension may not exist. In section 3, we will look at some examples, in order to understand why this is so. What the examples suggest, and what we will see in section 4, is that the Hermitian form can be defined only on appropriately \small" representations in each infinitesimal equivalence class. In the example of the various function spaces on X, compactly supported smooth functions are appropriately small, and will often carry an invariant Hermitian form. Distributions, on the other hand, are generally too large a space to admit an invariant Hermitian form. Here is a precise statement. (We will write V ∗ for the space of continuous linear functionals on V , endowed with the strong topology (see section 8).) Theorem 1.5 (Casselman, Wallach, and Schmid; see [CasG], [SchG], and section 4). Suppose (π; V ) is an admissible irreducible representation of a reductive Lie group G on a reflexive Banach space V . Define (π!; V !) = analytic vectors in V ; (π1; V 1) = smooth vectors in V ; 4 DAVID A. VOGAN, JR. (π−∞; V −∞) = distribution vectors in V = dual of (V 0)1, and (π−!; V −!) = hyperfunction vectors in V = dual of (V 0)!. Each of these four representations is a smooth representation of G in the infinites- imal equivalence class of π, and each depends only on that equivalence class. The inclusions V ! ⊂ V 1 ⊂ V ⊂ V −∞ ⊂ V −! are continuous, with dense image. Any invariant Hermitian form h; iK on VK extends uniquely to continuous G- ! 1 invariant Hermitian forms h; i! and h; i1 on V and V . The assertions about Hermitian forms will be proven in Theorem 9.16. The four representations appearing in Theorem 1.5 are called the minimal glob- alization, the smooth globalization, the distribution globalization, and the maximal globalization respectively. Unless π is finite-dimensional (so that all of the spaces in the theorem are the same) the Hermitian form will not extend continuously to the distribution or maximal globalizations V −∞ and V −!. We will be concerned here mostly with representations of G constructed using complex analysis, on spaces of holomorphic sections of vector bundles and general- izations. In order to use these constructions to get unitary representations, we need to do the analysis in such a way as to get the minimal or smooth globalizations; this will ensure that the Hermitian forms we seek will be defined on the representa- tions. A theorem of Hon-Wai Wong (see [Wng] or Theorem 7.21 below) says that Dolbeault cohomology leads to the maximal globalizations in great generality. This means that there is no possibility of finding invariant Hermitian forms on these Dolbeault cohomology representations except in the finite-dimensional case. We therefore need a way to modify the Dolbeault cohomology construction to produce minimal globalizations rather than maximal ones. Essentially we will follow ideas of Serre from [Ser], arriving at realization of minimal globalization represen- tations first obtained by Tim Bratten in [Br].